Effect of Structural Heterogeneity in Chemical Composition on Online

Mar 16, 2017 - Effect of Structural Heterogeneity in Chemical Composition on Online Single-Particle Mass Spectrometry Analysis of Sea Spray Aerosol Pa...
0 downloads 10 Views 998KB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

The Effect of Structural Heterogeneity in Chemical Composition on Online Single Particle Mass Spectrometry Analysis of Sea Spray Aerosol Particles Camille M Sultana, Douglas B. Collins, and Kimberly Ann Prather Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b06399 • Publication Date (Web): 16 Mar 2017 Downloaded from http://pubs.acs.org on March 22, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30

Environmental Science & Technology

The Effect of Structural Heterogeneity in Chemical Composition on Online Single Particle Mass Spectrometry Analysis of Sea Spray Aerosol Particles Camille M. Sultana1, Douglas B. Collins1†, Kimberly A. Prather1,2,* 1

Department of Chemistry and Biochemistry, University of California, San Diego, La

Jolla, CA 92093-0314; 2

Scripps Institution of Oceanography, University of California, San Diego, La Jolla, CA

92093; 1

ABSTRACT

2

Knowledge of the surface composition of sea spray aerosols (SSA) is critical for

3

understanding and predicting climate-relevant impacts. Offline microscopy and

4

spectroscopy studies have shown dry supermicron SSA tend to be spatially

5

heterogeneous particles with sodium and chloride rich cores surrounded by organic

6

enriched surface layers containing minor inorganic seawater components such as

7

magnesium and calcium. At the same time, single particle mass spectrometry reveals

8

several different mass spectral ion patterns, suggesting that there may be a number of

9

chemically distinct particle types. This study investigates factors controlling single

10

particle mass spectra of nascent supermicron SSA. Depth profiling experiments

11

conducted on SSA generated by a fritted bubbler and total ion intensity analysis of SSA

12

generated by a marine aerosol reference tank were compared with observations of

13

ambient SSA observed at two coastal locations. Analysis of SSA produced utilizing

14

controlled laboratory methods reveals that single particle mass spectra with weak sodium

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 30

15

ion signals can be produced by the desorption of the surface of typical dry SSA particles

16

comprised of salt cores and organic rich coatings. Thus, this lab-based study for the first

17

time unifies findings from offline and online measurements as well as lab and field

18

studies of SSA particle mixing state.

19

1. INTRODUCTION

20

Atmospheric aerosols impact climate by interacting directly with incoming solar radiation

21

and indirectly through influencing cloud properties by serving as cloud condensation

22

(CCN) and ice nuclei.1,2 Sea spray aerosols (SSA) represent one of the most abundant

23

types of tropospheric aerosols.3 The ability to determine the radiative forcing of

24

anthropogenic aerosols is limited by the uncertainty of the impact of large natural sources

25

of aerosols on current total aerosol radiative forcing.4 SSA are ejected into the

26

atmosphere when bubbles burst at the air-sea interface.5 Previous field and lab studies

27

have shown that SSA is a complex mixture of inorganic salts and an array of dissolved

28

and particulate organic components.6,7 Modeling studies have indicated that the mixing

29

state of submicron SSA can have important effects on predicted CCN concentrations,8–11

30

while the three dimensional chemical structure and mixing state of supermicron SSA can

31

affect light scattering due to changes in water uptake at sub-saturated relative

32

humidities.12,13 Additionally, the supermicron size range is where the bulk of SSA surface

33

area resides, and thus plays a key role in light scattering and interactions with gaseous

34

species.14–16

35

In an effort to characterize the diversity in chemical mixing state of SSA, field

36

studies employing off-line microscopy and spectroscopy techniques have helped illustrate

37

the array of submicron SSA particle types.7,17–21 In contrast, offline chemical analysis

ACS Paragon Plus Environment

2

Page 3 of 30

Environmental Science & Technology

38

techniques have found the vast majority of dry supermicron particles to be phase-

39

separated inorganic cuboids with amorphous coatings.7,19–32 Elemental analyses have

40

shown that the inorganic cores are mainly composed of sodium chloride, while organic

41

compounds along with minor inorganic species, such as magnesium, potassium, and

42

calcium, are heavily enriched in the amorphous coating or surface localized nodules.19–26

43

These results are in agreement with efflorescence studies of particles generated from

44

natural seawater and model salt solutions, which demonstrated that after drying the

45

particles consist of sodium chloride cores with the particle surface enriched in the minor

46

inorganic components such as magnesium, potassium, and calcium.33–40 Although the

47

amount of the coating material can vary between particles, offline chemical analysis has

48

generally shown a single type of dry supermicron SSA particle with the population

49

represented as an internal mixture of salt and organic species. While it is known that the

50

bubble bursting process can eject both fragmented and intact microbiological cells,23,41

51

findings from spectromicroscopy techniques suggest that particles containing whole cells

52

are rare. However, the collection and analysis methods are often not optimal for

53

preserving cellular structures, and the sample throughput is often low. Consistent with

54

this result, single bubble bursting experiments have reported the fraction of collected film

55

and jet drops containing bacteria cells to be ~1% and less than 0.1%, respectively.41–43

56

However, analysis of dried SSA by online single particle mass spectrometry (SPMS) has

57

described an externally mixed population. Analysis based upon mass spectral signatures

58

has identified three distinct particle types contributing significantly to the supermicron

59

SSA population: sea salt (SS), sea salt with organic carbon (SSOC), and biological or

60

magnesium (Bio) particles.28,44–46 Bridging the data from both offline and online methods,

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 30

61

as well as understanding the factors that affect the mixing state of supermicron SSA

62

particles is a requirement for understanding the water uptake and light scattering

63

properties of marine aerosols.

64

Particles with core-shell morphologies, comprised of specific components making up the

65

core of the particle with other chemical species forming an outer layer were generated in

66

previous laboratory studies and studied with SPMS.47–51 By varying the energy of the

67

laser pulse that desorbs and ionizes the core-shell particles, these studies have illustrated

68

that depth profiling of particles can be accomplished in real time. Lower laser fluence

69

tends to partially desorb the particle, generating mass spectra (MS) that are representative

70

of the surface components, whereas higher laser fluence typically desorbs a greater

71

fraction of the particle, generating MS with greater ion signal contribution from core

72

components. However, it has also been shown that even performing laser

73

desorption/ionization using a consistent laser pulse energy on well-controlled particles of

74

similar composition, there can be considerable variation in the laser fluence experienced

75

by individual particles resulting in a range of mass spectra.47,49,51,52 Considering the

76

structural observations of SSA by spectromicroscopy, combined with the variations in

77

desorption/ionization common to SPMS, the range of MS patterns generated by SSA may

78

be, at least in part, a result of the structural heterogeneity of dry supermicron SSA. Total

79

ion intensity of each mass spectrum has been utilized as an indirect measure of the extent

80

of particle desorption (proportional to the laser fluence experienced by the particle), and

81

on average, mass spectra of SSA particles that had low total ion intensity were relatively

82

rich in Mg, K, and Ca.22

ACS Paragon Plus Environment

4

Page 5 of 30

83

Environmental Science & Technology

This study extends the complexity of previous single particle mass spectrometry

84

studies in an effort to inform the existing descriptions of the chemical mixing state of

85

supermicron SSA particles. Aerosol time-of-flight mass spectrometry (ATOFMS) was

86

used to perform a detailed analysis of laser desorption/ionization mass spectra of dry

87

supermicron SSA sampled from three different sources: (1) SSA generated by bubbling

88

seawater using a sintered frit, (2) SSA generated in a Marine Aerosol Reference Tank

89

(MART), and (3) SSA sampled in ambient aerosol during two coastal field studies.

90

Depth profiling of frit-generated SSA was performed using ATOFMS. The

91

results from the depth profiling study, the first to be performed on SSA particles utilizing

92

SPMS, are extended to the MART and ambient generation schemes by examining the

93

relationship of the relative sodium ion signal to mass spectral patterns and total ion

94

intensity. Finally, this new approach is applied to sodium deficient and sodium rich SSA

95

mass spectra collected during two coastal field studies. This study aims to unify the

96

description of the mixing state and chemical composition of supermicron SSA by offline

97

spectroscopy and microscopy techniques with real-time analysis by single particle mass

98

spectrometry.

99

2. METHODS

100

2.1 SSA generation

101

2.1.1 Fritted Bubbler

102

On 1/18/16 seawater was collected from the coastal Pacific Ocean at Scripps Pier

103

(La Jolla, CA; 32°51´56.8"N: 117° 15´38.48"W; 275 m offshore) at least 5 meters below

104

the low tide line and passed through a sand bed filter to remove large debris. The water

105

was then filtered through 0.8 µm and then 0.2 µm filters to remove biological particles

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 30

106

such as bacteria and phytoplankton cells. 350 mL of the filtered seawater was then added

107

to a 500 mL gas washing bottle or ‘fritted bubbler’ (Ace Glass) previously combusted at

108

400 °C for 6 h. Zero air (Sabio 1001) was flowed through the sintered glass filter or ‘frit’

109

(Pore Size C: 25-50 µm), submerged at a depth of roughly 17 cm, at a rate of 0.04 L/min

110

to generate bubbles that rose through the seawater column, breaking at the surface to

111

generate SSA particles used in the analyses described below. Additional zero air was

112

added to the sample line after the bubbler to maintain a flow rate above 1 LPM (~1.25

113

LPM). Silica gel diffusion dryers were placed between the bubbler and the ATOFMS

114

inlet to dry the particles (RH < 10%).

115

2.1.2 MART

116

Coastal Pacific seawater (60 L) was collected from the ocean surface at Scripps

117

Pier (La Jolla, CA; 32°51´56.8"N: 117° 15´38.48"W; 275 m offshore) on 6/11/13 16:00

118

and added without treatment or filtering to a 100 L MART system.53 At the time of

119

collection, the chlorophyll-a concentration, water temperature, and salinity were 0.04

120

mg/m3, 19.4 °C, and 33.6 PSU. SSA particles were generated in the MART using the

121

pulsed plunging waterfall technique described in detail previously, with a 4 second

122

waterfall duty cycle.53 Two silica gel diffusion dryers were placed between the MART

123

and the ATOFMS inlet to dry the particles (RH ~15%).

124

2.1.3 CIFEX

125

ATOFMS measurements of ambient aerosols were made during the Cloud

126

Indirect Forcing Experiment (CIFEX) at a coastal site in Trinidad Head, CA (41.05° N:

127

124.15° W) in April 2004.54 Aerosols were collected at a height of 10 m, and the

128

sampling inlet was heated to maintain the RH at 55%.

ACS Paragon Plus Environment

6

Page 7 of 30

Environmental Science & Technology

129

2.1.3 CalWater2

130

ATOFMS measurements of ambient aerosols were made during the CalWater2

131

field study at a coastal site in Bodega Bay, CA on the grounds of Bodega Marine

132

Laboratory (39.32° N: 123.07° W) from January to March 2015.55 Aerosols were

133

collected at a height of five meters and silica gel diffusion dryers were placed before the

134

ATOFMS inlet to reduce the RH of sampled air to ~15%.

135

2.2 Measurement of SSA Composition via ATOFMS

136

The size-resolved chemical composition of individual dry SSA particles with

137

vacuum aerodynamic diameters (dva) between 1 - 3 µm were measured in real time by

138

ATOFMS. The measured vacuum aerodynamic size distribution of the sampled particles

139

is provided in Figure S1 along with estimated aerodynamic and volume equivalent

140

distributions (Method S1). A detailed discussion of ATOFMS has been given

141

previously,56,57 so only a brief description follows here. Aerosol particles are drawn into a

142

differentially pumped vacuum chamber through a converging nozzle inlet, wherein

143

particles are accelerated to their size-dependent terminal velocities. Aerosol particles are

144

sized based on the time required to transit two continuous wave laser beams (532 nm).

145

The velocity is converted to vacuum aerodynamic diameter via calibration with

146

polystyrene latex spheres (Invitrogen) of known diameter and density. When each

147

particle arrives in the ion source region of the mass spectrometer, a Q-switched Nd:YAG

148

laser pulse (266 nm wavelength, 8 ns pulse width, 700 µm spot size), triggered based on

149

the velocity of particle, desorbs and ionizes each particle’s chemical components. Laser

150

pulse energy was kept constant at approximately 1.1-1.3 mJ for the MART, CIFEX, and

151

CalWater2 studies. For the frit-generated SSA studies, the energy of the laser pulse was

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 30

152

varied between 0.6-1.5 mJ, and mass spectra from 800-1300 particles were collected at

153

each laser energy. Below 0.6 mJ, the percentage of particles generating MS dropped

154

considerably, precluding the use of such low laser pulse energies for these experiments.

155

The exact laser fluence that each particle experiences is variable even at a constant laser

156

energy setting due to the Gaussian profile of the laser beam, hot spots in the laser beam,

157

and shot-to-shot variations in laser pulse characteristics.52 The positive and negative ions

158

produced are detected by a dual-polarity reflectron time-of-flight mass spectrometer.

159

Single particle mass spectra and size data were analyzed using the software toolkit,

160

FATES.58

161

2.3 Analysis of Particle Mass Spectra

162

Note that all analyses included in this study were performed on particles with dva

163

between 1 and 3 µm. Particles were binned by the fraction of positive mass spectral

164

intensity attributable to sodium and sodium chloride containing ions (FNa): 23Na+, 46Na2+,

165

81,83

166

correspond to the most likely ion produced at a specific mass-to-charge ratio (m/z). For

167

consistency, only positive sodium ion markers were utilized as negative MS were not

168

detected for every instance of a positive MS. Note that total positive ion intensities are

169

the summation of all positive ion signal and are normalized to the maximum value within

170

each dataset. Based upon similarity to the mass spectral fingerprints of the lab-generated

171

SSA a subset of mass spectra most likely to represent freshly generated SSA particles

172

were identified within the field study datasets. Mass spectra that contained indications of

173

terrestrial or anthropogenic sources (e.g. nitrate, biomass burning, dust, soot) were

174

eliminated. Mass spectra identified as likely generated by cellular material based upon

Na2Cl+, 139,141,143Na3Cl2+, 197,199,201Na4Cl3+, 255,257,259Na4Cl3+. Peak assignments

ACS Paragon Plus Environment

8

Page 9 of 30

Environmental Science & Technology

175

comparison to the literature, with dominant potassium and phosphate ion markers,59–62

176

were not included in the SSA analysis for the field studies due to the possibility of a

177

biological terrestrial source. See supporting information for more detail on the analysis of

178

the laboratory and field data.

179

3. RESULTS AND DISCUSSION

180

3.1 Effect of Laser Power on Supermicron SSA Mass Spectra

181

Supermicron SSA particles (dry dva = 1–3 µm) generated from natural seawater

182

utilizing a fritted bubbler were analyzed via ATOFMS, varying the pulse energy of the

183

desorption/ionization laser from 0.6 to 1.5 mJ. It is important to note that the bubbled

184

seawater was filtered to remove most insoluble biological components such as bacteria

185

and phytoplankton cells and preclude such material from being aerosolized and

186

contributing to the mass spectra generated. Data from all laser energy conditions were

187

compiled, and then the average normalized MS for particles with differing fractions of

188

positive mass spectral intensity from sodium were calculated. For the purposes of this

189

study, all particles with FNa, fraction of sodium-containing positive ion signal, greater and

190

less than 0.4 are referred to as Na-Rich and Na-Deficient, respectively. Mass spectra with

191

a lower FNa have higher relative contributions from components that are expected to exist

192

at the surface of dried SSA particles, including calcium (40Ca+, 57CaOH+, 75CaCl+

193

96

194

109,111,113

195

59

196

S2). Interestingly, the average normalized MS with high (0.8-1), medium (0.4-0.6), and

197

low (0.0-0.1) FNa are similar to the representative mass spectral fingerprints of the SS,

Ca2O+, 145,147,149CaCl3-), magnesium (24Mg+, 129,131,133MgCl3-), potassium (39K+, KCl2-), and organic material (27C2H3+, 27CHN+, 37C3H+, 43C2H3O+, 43CHNO+,

C3H9N+, 26CN-, 42CNO-, 43C2H3O-) in both negative and positive mass spectra (Figure

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 30

198

SSOC, and Bio type particles that have been identified by previous SPMS analyses.44,46,63

199

Based solely on these results two potential explanations present themselves. Firstly the

200

three types of mass spectra could be derived from a single population of phase-separated

201

particles with chemical spatial heterogeneity, such as the core-shell morphology

202

previously described. Alternatively a collection of three somewhat distinct particle types

203

could exist in the aerosol population. Results from the different laser pulse energy trials

204

were used to distinguish between these two possibilities by providing information on the

205

three dimensional chemical morphology of the particles.

206

Previous studies utilizing SPMSs have reduced the pulse energy of the

207

desorption/ionization laser to selectively detect components on the surface of particles

208

with known core-shell morphologies.47–51 As the laser pulse energy was increased from

209

0.6 to 1.5 mJ the fraction of Na-Rich MS dramatically increased from 39% to 77.5%

210

(Figure 1). As indicated in the schematic in Figure 1, the greater sodium signal at higher

211

laser energy (more complete particle desorption/ionization) is consistent with structurally

212

heterogeneous particles in which sodium is located in the core. Lower laser energies were

213

more likely to desorb/ionize only small amounts of each particle, and the generated mass

214

spectra had relatively greater contributions from components such as magnesium and

215

organics. As discussed earlier, spectromicroscopy studies have shown that dried nascent

216

SSA are dominated by a single type of structurally heterogeneous particles with

217

magnesium, calcium, sulfate, potassium, and organics concentrated on the outside in an

218

amorphous coating or sometimes in distinct nodules surrounding cores of sodium

219

chloride.20–23 The depth profiling results indicate that the variety of supermicron SSA

ACS Paragon Plus Environment

10

Page 11 of 30

Environmental Science & Technology

220

mass spectra observed by ATOMFS could be attributed to a population of core-shell

221

particles encountering a range of desorption/ionization conditions within the instrument.

222

3.2 Relationship of Chemical Signal to Total Positive Spectrum Ion Intensity

223

Studies of lab-generated particles with well characterized and controlled core-

224

shell morphologies have shown that total positive ion intensity increases with increasing

225

relative contribution from components in the core of the particle.47,64 The use of total ion

226

intensity as a method to discriminate between mass spectra that were generated in

227

association with varying degrees of particle desorption (surface vs. core) has been

228

employed previously for lab-generated SSA.22 Particles analyzed in this study, however,

229

were grouped by the fraction of positive ion signal from sodium rather than by the

230

explicit total ion intensity as in the prior study. The total positive ion intensity

231

distribution for each group of mass spectra binned by FNa was calculated for all

232

supermicron particles. The results for the 1.2 mJ laser energy trial are shown in Figure 2,

233

and are representative of the results across all laser pulse energy (0.6-1.5 mJ) ranges

234

examined (Figure S3). At higher FNa the total positive ion intensity distribution shifted to

235

higher values (Figure 2-3). The increase in total positive ion intensities with increasing

236

relative sodium ion signal is consistent with increasing particle desorption as shown in

237

the conceptual cartoon in Figure 1. These results are in good agreement with previous

238

analyses of SSA particles by ATOFMS that grouped particles explicitly by total mass

239

spectral intensities.22 However, even when sodium rich mass spectra were detected,

240

particle desorption was likely not complete as the total positive ion MS intensity did not

241

increase with size (Figure S4).

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 30

242

The generally low positive ion intensity of Na-Deficient mass spectra supports the

243

conclusion that these MS are mainly representative of the particle surface which has been

244

shown to be enriched in organics and minor inorganic seawater components via offline

245

techniques. However, it appears that the outlying sodium deficient mass spectra with

246

relatively high total ion intensities also have high contributions from positive calcium ion

247

markers (40Ca+, 57CaOH+, 75,77CaCl+, 96Ca2O+) (Figure 3). This behavior is likely still

248

attributable to the morphology of dry SSA particles. For both SSA generated by the

249

fritted bubbler and the MART the relative calcium ion contribution and the relative

250

sulfate ion contribution (48SO-, 64SO2-, 80SO3-) have a positive correlation (r2bubbler = 0.35,

251

r2MART = 0.44) (Figure S5). Calcium sulfate, calcium sulfate hydrates, and sodium

252

calcium sulfate salts are well known to precipitate and crystallize initially during the

253

dehydration of seawater and seawater droplets.36,37,65 Nodules or crystals rich in calcium

254

and sulfur have been observed in dry SSA particles adhered to the outside of sodium

255

chloride rich cores.20,22 It is possible that desorption and ionization of calcium sulfate

256

crystals on the surface of dry SSA particles generates sodium deficient mass spectra but

257

with more ion intensity than when only desorption of the amorphous and relatively thin

258

coating occurs. The infrequency of these relatively intense calcium rich mass spectra

259

may be because these crystals or nodules are less ubiquitous than the amorphous coating

260

and localized to a specific region of a particle. Overall the relationship between the

261

positive mass spectral ion markers and total positive ion intensity is consistent with the

262

known core-shell morphologies of dry supermicron SSA.

263

3.3 SSA Mass Spectral Consistency between Laboratory and Field Studies

ACS Paragon Plus Environment

12

Page 13 of 30

264

Environmental Science & Technology

The data analyses described in the previous sections for SSA generated from

265

bubbled seawater were also performed on dry supermicron SSA generated by a MART

266

and from two field studies (CIFEX, CalWater2). For the MART and field studies, the

267

particles were binned by FNa (Table S1) and the average mass spectra were similar to the

268

results from SSA generated by bubbled seawater (Figure S2). In addition, for all four

269

datasets the population of mass spectra with the highest FNa (0.8-1) had higher total ion

270

intensities than the preponderance of mass spectra with the lowest FNa (0-0.1). These

271

results strongly suggest that the supermicron SSA generated by the fritted bubbler,

272

MART, and sampled during field studies are all similar in composition and morphology.

273

Therefore, for all datasets the supermicron SSA population is likely dominated by a

274

single particle type that, once dried, is structurally heterogeneous. Such particles exhibit a

275

sodium chloride core with a surface enriched in organic material and minor inorganic

276

seawater components. Some differences in the thickness of the sodium chloride deficient

277

particle coating are likely to exist, as illustrated by images of typical ambient, MART,

278

and frit generated dry SSA particles from prior studies (Figure S6).28 While the

279

morphology of dried SSA particles will not be representative of their hydrated structure

280

under relatively humid ambient conditions, studying dried phase-separated particles

281

provides a standard and experimentally simple common basis for analyzing chemical

282

composition across an array of different methods.

283

Results from this study demonstrate that sodium deficient mass spectra, which are

284

relatively rich in species such as magnesium, can be generated from desorption and

285

ionization of the surface of dried SSA. Since the acquisition of mass spectra that

286

differentially represent the core and surface coating of dry SSA is mostly a function of

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 30

287

the instrumental operating conditions (e.g., laser pulse energy), the ratio of sodium rich to

288

sodium deficient mass spectra should be relatively constant. To examine this hypothesis,

289

the fraction of identified fresh SSA Na-Rich and Na-Deficient mass spectra within two-

290

hour time bins was calculated for two field studies performed on the west coast of North

291

America, CIFEX and CalWater2 (Figure 4). In both studies the fraction of Na-Deficient

292

mass spectra to the fraction of Na-Rich mass spectra shows a generally positive trend as

293

expected, though the correlations are poor (r2CIFEX = 0.28, r2CW2 = 0.17). Notably, the

294

CIFEX results here are in close agreement with a previous analysis where the MS were

295

grouped by a clustering algorithm and “fresh SS” and “Mg-type” particle types, very

296

similar to the average MS for high and low FNa (Figure S2) respectively, were found to be

297

well correlated.63

298

To further support the assignment of the Na-Deficient mass spectra as fresh SSA,

299

versus terrestrial or anthropogenic particle types, the mean relative nitrate and sulfate ion

300

signal (46NO2-, 62NO3-, 97HSO4-, 125H(NO3)2-, 147Na(NO3)2-, 160HNO3HSO4-,

301

188

302

phase nitric and sulfuric acid is indicative of air masses influenced by anthropogenic

303

pollution and secondary aging processes. It is anticipated that the fraction of Na-

304

Deficient, fresh SSA mass spectra will be negatively correlated with air masses not

305

abundant in fresh SSA, and therefore the mean relative nitrate ion signal and sulfate ion

306

signal. The results show a general negative trend (r2CIFEX = 0.48, r2CW2 = 0.35).

(HNO3)2NO3-) for all mass spectra detected was calculated for each time bin. Particle

307

While the resulting relationships generally support the conclusion that the

308

identified Na-Rich and Na-Deficient mass spectra originate from the same particle type,

309

large variation exists in the data. Variations in the Na-Rich to Na-Deficient ratio may be

ACS Paragon Plus Environment

14

Page 15 of 30

Environmental Science & Technology

310

driven by the analytical difficulties associated with utilizing field study datasets for this

311

specific type of investigation. This analysis depends upon accurately identifying SSA

312

mass spectra in an uncontrolled environment. However, there are a variety of particle

313

types, such as dust, biomass burning, and cellular material, which can generate mass

314

spectra rich in magnesium, calcium, potassium, and/or organic ion markers. It is possible

315

that some Na-Deficient mass spectra included were not generated by surface desorption

316

of SSA particles. Additionally it is key to include only MS from fresh SSA that have not

317

been aged by secondary processes which have been shown to change the three

318

dimensional physicochemical structure of SSA.22 However, negative MS, which

319

frequently contain the secondary processing ion markers, are generated with less

320

frequency than positive MS, particularly for low ion intensity mass spectra.

321

In addition, changes in seawater chemistry have been shown to influence SSA

322

organic content and therefore also likely modify the thickness of the sodium deficient

323

shell.10,66,67 It is expected that increases in the thickness of the organic coating on dry

324

supermicron SSA particles would directly correspond to an increase in the fraction of Na-

325

Deficient mass spectra generated at a constant laser pulse energy.47,50 Therefore some of

326

the variation in the ratio of Na-Rich to Na-Deficient MS noted in the field studies may be

327

a result of changing seawater chemistry. While the paradigm that this study establishes

328

with respect to interpretation of single particle mass spectra of SSA should find facile

329

utility for studies of nascent SSA, there are still difficulties in applying this framework to

330

complex field study datasets.

331

3.4 Contribution of Particulate Biological Components to SSA MS

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 30

332

The MART and bubbled seawater datasets were highly similar, however only the

333

bubbled seawater was filtered to remove microbial cells. This suggests that intact cells or

334

cell fragments were not abundant in the MART supermicron SSA or that mass spectra

335

generated by aerosolized cells could not be reliably distinguished from those obtained

336

from other salt and organic matter-containing supermicron SSA particles. As described,

337

ion signatures which have been previously associated with cellular single particle mass

338

spectra, such as organic, potassium, calcium, and phosphate ion markers, can be

339

generated from the surface desorption of the structurally heterogeneous dried

340

supermicron particles. It is important to note that while intact biological particles can be

341

ejected in and have been identified in SSA,23,68 it is difficult to unequivocally assign a

342

mass spectral pattern to aerosolized marine cells without more subtle data analysis

343

methods or further information about the particles’ properties. Therefore, further

344

correlated information (e.g., fluorescence) may be required to identify and quantify the

345

abundance of microbial cells ejected from the surface ocean.

346

3.5 Interpreting mass spectra from SPMS studies of SSA

347

Other than ejected particles composed of particulate biological components, the

348

chemical composition of most supermicron SSA in the size regime examined (dva = 1-3

349

µm) is expected to be relatively similar between particles.7 However within each particle,

350

spectromicroscopy analyses have shown a great deal of spatial heterogeneity in the

351

structure of dry SSA particles.19–24,26,69 Studies of dehydrated natural and model seawater

352

droplets show that particles exhibit a core-shell morphology, with sodium chloride rich

353

cores and nodules of calcium sulfate and an amorphous coating rich in magnesium,

354

potassium, organics and other minor inorganic components at the surface.33–40 SPMS

ACS Paragon Plus Environment

16

Page 17 of 30

Environmental Science & Technology

355

analysis of morphologically-controlled core-shell particles has shown that even at a

356

constant laser pulse energy setting there is wide particle-to-particle variability in the ratio

357

of signal from surface and core components in single particle mass spectra.47,49,51 This

358

variation exists despite the chemical and morphological similarities between the lab-

359

generated core-shell particles. As seen in this study, mass spectra with a wide range of

360

sodium fractions were generated for all laser energies, which is consistent with previous

361

studies of core-shell particles. However, the wide variability in signal response illustrates

362

the complexity in analyzing SPMS data sets.

363

The schematic in Figure 5 illustrates that when analyzing single particle mass

364

spectra of SSA, or any population of particles with structural heterogeneity, distinctions

365

between relative mass spectra are not necessarily indicative of an externally mixed,

366

chemically distinct particle population. Rather we have shown that varying degrees of

367

laser desorption and ionization of core-shell SSA particles can produce MS with high,

368

medium, and low sodium ion contributions which have been previously characterized as

369

originating from distinct SS, SSOC, and Bio SSA particle types respectively.28,44–46

370

Utilizing the paradigm established, increases in the fraction of low intensity and sodium

371

deficient mass spectra generated by dried supermicron SSA should likely be interpreted

372

as increases in the thickness of the amorphous coating on typical SSA particles rather

373

than an appearance of a new distinct sodium deficient particle type. Leveraging depth

374

profiling analysis and examining trends in total ion intensity coupled with known

375

morphology from spectroscopy and microscopy studies has helped enable this distinction.

376

These results bring real-time chemical analysis of SSA by SPMS into agreement with

377

offline microscopy and spectroscopy descriptions of the supermicron SSA population,

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 30

378

which dictates SSA’s direct light scattering properties and interactions with gaseous

379

species.14–16 Importantly, this analysis and new understanding can serve as a foundation

380

for SPMS analysis of SSA, which could be extended in the future to the more complex

381

submicron SSA population. It is important to note that the explicitly varying the laser

382

pulse energy proved to be a useful tool for understanding the three-dimensional structural

383

morphology and composition of atmospheric SSA particles.

384

ASSOCIATED CONTENT

385

Supporting Information Method for converting particle diameter; Experimental method

386

for generating and imaging SSA particles in Figure S6; A table providing the number of

387

mass spectra in each sodium fraction bin for all experiments; Size distribution for

388

particles detected and analyzed by the ATOFMS; Average relative mass spectra are

389

shown for supermicron SSA and grouped by the fraction of positive mass spectra

390

intensity from sodium; The total positive ion spectrum intensity versus size for the

391

bubbled seawater experiment; The total ion intensity distribution for all bubbled seawater

392

data; The calcium ion fraction plotted against the sulfate ion fraction for individual mass

393

spectra; Compilation of images of SSA particles.

394

AUTHOR INFORMATION

395

Corresponding Author

396

*E-mail: [email protected]. Phone: +1 858 822 5312. Address: Department of

397

Chemistry and Biochemistry, University of California, San Diego, La Jolla, CA 92093-

398

0314.

399

Present Addresses

400

† Department of Chemistry, University of Toronto, Toronto, ON, Canada M5S 3H6

ACS Paragon Plus Environment

18

Page 19 of 30

Environmental Science & Technology

401

ACKNOWLEDGMENTS

402

This work was supported by the National Science Foundation through the Centers of

403

Chemical Innovation Program via the Center for Aerosol Impacts on Climate and the

404

Environment (CHE-1305427).

405

supplemental atomic force microscopy analysis, all collaborators involved in the MART

406

microcosm studies, notably C. Lee, as well as J. Holecek and G. Cornwell for field study

407

data collection. In addition the authors would like to acknowledge the Bodega Marine

408

Reserve, University of California Davis, and UC Natural Reserve System for use of

409

Bodega Marine Laboratory facilities.

410

REFERENCES

411 412

(1)

Haywood, J.; Boucher, O. Estimates of the direct and indirect radiative forcing due to tropospheric aerosols: A review. Rev. Geophys. 2000, 38 (4), 513–543.

413 414

(2)

Lohmann, U.; Feichter, H. Global indirect aerosol effects: a review. Atmos. Chem. Phys. 2005, 5, 715–737.

415 416 417 418

(3)

Textor, C.; Schulz, M.; Guibert, S.; Kinne, S.; Balkanski, Y.; Bauer, S.; Berntsen, T.; Berglen, T.; Boucher, O.; Chin, M.; et al. Analysis and quantification of the diversities of aerosol life cycles within AeroCom. Atmos. Chem. Phys. 2006, 6, 1777–1813.

419 420 421

(4)

Stocker, T. F.; Qin, D.; Plattner, G. K.; Tignor, M.; Allen, S. K.; Boschung, J.; Nauels, A.; Xia, Y.; Bex, V.; Midgley, P. M. IPCC, 2013: Summary for Policymakers. IPCC WG1 5th Assessment Report.; 2013.

422 423

(5)

Blanchard, D. C.; Woodcock, A. H. Bubble Formation and Modification in the Sea and its Meteorological Significance. Tellus IX 1957, 9 (2), 145–158.

424 425 426

(6)

Quinn, P. K.; Collins, D. B.; Grassian, V. H.; Prather, K. A.; Bates, T. S. Chemistry and Related Properties of Freshly Emitted Sea Spray Aerosol. Chem. Rev. 2015, 115 (4383–4399), 150406123611007.

427 428

(7)

Gantt, B.; Meskhidze, N. The physical and chemical characteristics of marine primary organic aerosol: a review. Atmos. Chem. Phys. 2013, 13 (8), 3979–3996.

429 430 431

(8)

Gantt, B.; Johnson, M. S.; Meskhidze, N.; Sciare, J.; Ovadnevaite, J.; Ceburnis, D.; O’Dowd, C. D. Model evaluation of marine primary organic aerosol emission schemes. Atmos. Chem. Phys. 2012, 12 (18), 8553–8566.

432 433 434

(9)

Meskhidze, N.; Xu, J.; Gantt, B.; Zhang, Y.; Nenes, a.; Ghan, S. J.; Liu, X.; Easter, R.; Zaveri, R. Global distribution and climate forcing of marine organic aerosol: 1. Model improvements and evaluation. Atmos. Chem. Phys. 2011, 11 (22), 11689–

The authors would like to thank O. Ryder for

ACS Paragon Plus Environment

19

Environmental Science & Technology

435

Page 20 of 30

11705.

436 437 438

(10)

O’Dowd, C. D.; Facchini, M. C.; Cavalli, F.; Ceburnis, D.; Mircea, M.; Decesari, S.; Fuzzi, S.; Yoon, Y. J.; Putaud, J.-P. Biogenically driven organic contribution to marine aerosol. Nature 2004, 431 (October), 676–680.

439 440 441

(11)

Westervelt, D. M.; Moore, R. H.; Nenes, a.; Adams, P. J. Effect of primary organic sea spray emissions on cloud condensation nuclei concentrations. Atmos. Chem. Phys. 2012, 12 (1), 89–101.

442 443 444

(12)

Saxena, P.; Hildemann, L. M.; McMurry, P. H.; Seinfeld, J. H. Organics alter hygroscopic behavior of atmospheric particles. J. Geophys. Res. 1995, 100 (D9), 18755.

445 446 447 448

(13)

Forestieri, S. D.; Cornwell, G. C.; Helgestad, T. M.; Moore, K. A.; Lee, C.; Novak, G. A.; Sultana, C. M.; Wang, X.; Bertram, T. H.; Prather, K. A.; et al. Linking variations in sea spray aerosol particle hygroscopicity to composition during two microcosm experiments. Atmos. Chem. Phys. 2016, 16 (14), 9003–9018.

449 450 451

(14)

Lewis, E. R.; Schwartz, S. E. Methods of Determining Size-Dependent Sea Salt Aerosol. In Sea Salt Aerosol Production: Mechanisms, Methods, Measurements and Models; 2004; pp 101–118.

452 453 454 455

(15)

Lewis, E. R.; Schwartz, S. E. Measurements and Models of Quantities Required to Evaluate Sea Salt Aerosol Production Fluxes. In Sea Salt Aerosol Production: Mechanisms, Methods, Measurements and Models-A Critical Review; 2013; pp 119–297.

456 457 458

(16)

Fan, T.; Toon, O. B. Modeling sea-salt aerosol in a coupled climate and sectional microphysical model: Mass, optical depth and number concentration. Atmos. Chem. Phys. 2011, 11 (9), 4587–4610.

459 460 461

(17)

Park, J. Y.; Lim, S.; Park, K. Mixing state of submicrometer sea spray particles enriched by insoluble species in bubble-bursting experiments. J. Atmos. Ocean. Technol. 2014, 31 (1), 93–104.

462 463 464

(18)

Orellana, M. V; Matrai, P. A.; Leck, C.; Rauschenberg, C. D.; Lee, A. M. Marine microgels as a source of cloud condensation nuclei in the high Arctic. Proc. Natl. Acad. Sci. U. S. A. 2011, 108 (33), 13612–13617.

465 466 467 468

(19)

Hawkins, L. N.; Russell, L. M. Polysaccharides , Proteins , and Phytoplankton Fragments : Four Chemically Distinct Types of Marine Primary Organic Aerosol Classified by Single Particle Spectromicroscopy. Adv. Meteorol. 2010, 2010, 612132.

469 470 471 472

(20)

Ault, A. P.; Moffet, R. C.; Baltrusaitis, J.; Collins, D. B.; Ruppel, M. J.; CuadraRodriguez, L. A.; Zhao, D.; Guasco, T. L.; Ebben, C. J.; Geiger, F. M.; et al. SizeDependent Changes in Sea Spray Aerosol Composition and Properties with Different Seawater Conditions. Environ. Sci. Technol. 2013, 47, 5603–5612.

473 474 475 476

(21)

Russell, L. M.; Hawkins, L. N.; Frossard, A. A.; Quinn, P. K.; Bates, T. S. Carbohydrate-like composition of submicron atmospheric particles and their production from ocean bubble bursting. Proc. Natl. Acad. Sci. 2010, 107 (15), 6652–6657.

ACS Paragon Plus Environment

20

Page 21 of 30

Environmental Science & Technology

477 478 479 480 481

(22)

Ault, A. P.; Guasco, T. L.; Ryder, O. S.; Baltrusaitis, J.; Cuadra-Rodriguez, L. a.; Collins, D. B.; Ruppel, M. J.; Bertram, T. H.; Prather, K. A.; Grassian, V. H. Inside versus outside: Ion redistribution in nitric acid reacted sea spray aerosol particles as determined by single particle analysis. J. Am. Chem. Soc. 2013, 135 (39), 14528–14531.

482 483 484 485

(23)

Patterson, J. P.; Collins, D. B.; Michaud, J. M.; Axson, J. L.; Sultana, C. M.; Moser, T.; Dommer, A. C.; Conner, J.; Grassian, V. H.; Stokes, M. D.; et al. Sea Spray Aerosol Structure and Composition Using Cryogenic Transmission Electron Microscopy. ACS Cent. Sci. 2016, 2, 40–47.

486 487 488 489 490

(24)

Ault, A. P.; Zhao, D.; Ebben, C. J.; Tauber, M. J.; Geiger, F. M.; Prather, K. A.; Grassian, V. H. Raman microspectroscopy and vibrational sum frequency generation spectroscopy as probes of the bulk and surface compositions of sizeresolved sea spray aerosol particles. Phys. Chem. Chem. Phys. 2013, 15 (207890), 6206–6214.

491 492 493 494

(25)

Chi, J. W.; Li, W. J.; Zhang, D. Z.; Zhang, J. C.; Lin, Y. T.; Shen, X. J.; Sun, J. Y.; Chen, J. M.; Zhang, X. Y.; Zhang, Y. M.; et al. Sea salt aerosols as a reactive surface for inorganic and organic acidic gases in the Arctic troposphere. Atmos. Chem. Phys. 2015, 15 (19), 11341–11353.

495 496 497

(26)

Posfai, M.; Anderson, J. R.; Buseck, P. R. Compositional variations of sea-saltmode aerosol particles from the North Atlantic. J. Geophys. Res. 1995, 100 (D11), 23063–23074.

498 499 500 501

(27)

Wise, M. E.; Freney, E. J.; Tyree, C. A.; Allen, J. O.; Martin, S. T.; Russell, L. M.; Buseck, P. R. Hygroscopic behavior and liquid-layer composition of aerosol particles generated from natural and artificial seawater. J. Geophys. Res. Atmos. 2009, 114 (3), 1–8.

502 503 504 505 506

(28)

Collins, D. B.; Zhao, D. F.; Ruppel, M. J.; Laskina, O.; Grandquist, J. R.; Modini, R. L.; Stokes, M. D.; Russell, L. M.; Bertram, T. H.; Grassian, V. H.; et al. Direct aerosol chemical composition measurements to evaluate the physicochemical differences between controlled sea spray aerosol generation schemes. Atmos. Meas. Tech. 2014, 7 (11), 3667–3683.

507 508 509 510

(29)

Lee, C.; Sultana, C. M.; Collins, D. B.; Santander, M. V; Axson, L.; Malfatti, F.; Cornwell, G. C.; Grandquist, J. R.; Deane, G. B.; Stokes, M. D.; et al. Advancing Model Systems for Fundamental Laboratory Studies of Sea Spray Aerosol using the Microbial Loop. J. Phys. Chem. A 2015.

511 512 513 514 515

(30)

Maskey, S.; Geng, H.; Song, Y. C.; Hwang, H.; Yoon, Y. J.; Ahn, K. H.; Ro, C. U. Single-particle characterization of summertime antarctic aerosols collected at King George island using quantitative energy-dispersive electron probe X-ray microanalysis and attenuated total reflection Fourier transform-infrared imaging techniques. Environ. Sci. Technol. 2011, 45 (15), 6275–6282.

516 517 518 519

(31)

Laskin, A.; Moffet, R. C.; Gilles, M. K.; Fast, J. D.; Zaveri, R. A.; Wang, B.; Nigge, P.; Shutthanandan, J. Tropospheric chemistry of internally mixed sea salt and organic particles: Surprising reactivity of NaCl with weak organic acids. J. Geophys. Res. Atmos. 2012, 117 (15), 1–12.

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 30

520 521

(32)

522 523 524 525

(33) Harmon, C. W.; Grimm, R. L.; McIntire, T. M.; Peterson, M. D.; Njegic, B.; Angel, V. M.; Alshawa, A.; Underwood, J. S.; Tobias, D. J.; Gerber, R. B.; et al. Hygroscopic growth and deliquescence of NaCl nanoparticles mixed with surfactant SDS. J. Phys. Chem. B 2010, 114 (7), 2435–2449.

526 527

(34)

Ge, Z. Z.; Wexler, A. S.; Johnston, M. V. Multicomponent aerosol crystallization. J. Colloid Interface Sci. 1996, 183 (1), 68–77.

528 529

(35)

Ge, Z.; Wexler, A. S.; Johnston, M. V. Deliquescence Behavior of Multicomponent Aerosols. J. Phys. Chem. A 1998, 5639 (97), 173–180.

530 531 532

(36)

Xiao, H. S.; Dong, J. L.; Wang, L. Y.; Zhao, L. J.; Wang, F.; Zhang, Y. H. Spatially resolved micro-raman observation on the phase separation of effloresced sea salt droplets. Environ. Sci. Technol. 2008, 42 (23), 8698–8702.

533 534 535

(37)

Tong, H. J.; Qian, Z. G.; Reid Jonathan, P.; Zhang, Y. H. High temporal and spatial resolution measurements of the rapid efflorescence of sea salt droplets. Wuli Huaxue Xuebao/ Acta Phys. - Chim. Sin. 2011, 27 (11), 2521–2527.

536 537 538 539

(38)

Gupta, D.; Eom, H. J.; Cho, H. R.; Ro, C. U. Hygroscopic behavior of NaClMgCl2 mixture particles as nascent sea-spray aerosol surrogates and observation of efflorescence during humidification. Atmos. Chem. Phys. 2015, 15 (19), 11273– 11290.

540 541 542

(39)

Cziczo, D. J.; Nowak, J. B.; Hu, J. H.; Abbatt, J. P. D. Infrared spectroscopy of model tropospheric aerosols as a function of relative humidity: Observation of deliquescence and crystallization. J. Geophys. Res. 1997, 102 (D15), 18843.

543 544 545 546

(40)

Liu, Y.; Yang, Z.; Desyaterik, Y.; Gassman, P. L.; Wang, H.; Laskin, A. Hygroscopic behavior of substrate-deposited particles studied by micro-FT-IR spectroscopy and complementary methods of particle analysis. Anal. Chem. 2008, 80 (3), 633–642.

547 548

(41)

Blanchard, D. C.; Syzdek, L. D. Water-to-air transfer and enrichment of bacteria in drops from bursting bubbles. Appl. Environ. Microbiol. 1982, 43 (5), 1001–1005.

549 550 551

(42)

Blanchard, D.; Syzdek, L. Bubble tube: apparatus for determining rate of collection of bacteria by an air bubble rising in water. Limnol. Oceanogr. 1974, 19 (1), 133–138.

552 553

(43)

Blanchard, D.; Syzdek, L. Seven problems in bubble and jet drop researches. Limnol. Oceanogr. 1978, 23 (3), 389–400.

554 555 556 557 558

(44)

Guasco, T. L.; Cuadra-Rodriguez, L. A.; Pedler, B. E.; Ault, A. P.; Collins, D. B.; Zhao, D.; Kim, M. J.; Ruppel, M. J.; Wilson, S. C.; Pomeroy, R. S.; et al. Transition Metal Associations with Primary Biological Particles in Sea Spray Aerosol Generated in a Wave Channel. Environ. Sci. Technol. 2014, 48 (2), 1324– 1333.

559 560 561

(45)

Gaston, C. J.; Furutani, H.; Guazzotti, S. a.; Coffee, K. R.; Bates, T. S.; Quinn, P. K.; Aluwihare, L. I.; Mitchell, B. G.; Prather, K. a. Unique ocean-derived particles serve as a proxy for changes in ocean chemistry. J. Geophys. Res. 2011, 116 (D18),

Leck, C.; Bigg, E. K. Comparison of sources and nature of the tropical aerosol with the summer high Arctic aerosol. Tellus 2008, 60B (1), 118–126.

ACS Paragon Plus Environment

22

Page 23 of 30

Environmental Science & Technology

562

D18310.

563 564 565 566

(46)

Prather, K. A.; Bertram, T. H.; Grassian, V. H.; Deane, G. B.; Stokes, M. D.; Demott, P. J.; Aluwihare, L. I.; Palenik, B. P.; Azam, F.; Seinfeld, J. H.; et al. Bringing the ocean into the laboratory to probe the chemical complexity of sea spray aerosol. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (19), 7550–7555.

567 568 569 570

(47)

Zelenyuk, A.; Juan, Y.; Chen, S.; Zaveri, R. a.; Imre, D. “Depth-profiling” and quantitative characterization of the size, composition, shape, density, and morphology of fine particles with SPLAT, a single-particle mass spectrometer. J. Phys. Chem. A 2008, 112 (4), 669–671.

571 572 573

(48)

Woods, E. I.; Smith, G. D.; Miller, R. E.; Baer, T. Depth Profiling of Heterogeneously Mixed Aerosol Particles Using Single-Particle Mass Spectrometry. Anal. Chem. 2002, 74 (7), 1642–1649.

574 575 576

(49)

Carson, P. G.; Johnston, M. V.; Wexler, A. S. Real-Time Monitoring of the Surface and Total Composition of Aerosol Particles. Aerosol Sci. Technol. 1997, 26 (4), 291–300.

577 578 579

(50)

Cai, Y.; Zelenyuk, A.; Imre, D. A High Resolution Study of the Effect of Morphology On the Mass Spectra of Single PSL Particles with Na-containing Layers and Nodules. Aerosol Sci. Technol. 2006, 40 (12), 1111–1122.

580 581 582

(51)

Cahill, J. F.; Fei, H.; Cohen, S. M.; Prather, K. a. Characterization of core–shell MOF particles by depth profiling experiments using on-line single particle mass spectrometry. Analyst 2015, 140 (5), 1510–1515.

583 584 585

(52)

Wenzel, R. J.; Prather, K. a. Improvements in ion signal reproducibility obtained using a homogeneous laser beam for on-line laser desorption/ionization of single particles. Rapid Commun. Mass Spectrom. 2004, 18 (13), 1525–1533.

586 587 588 589

(53)

Stokes, M. D.; Deane, G. B.; Prather, K.; Bertram, T. H.; Ruppel, M. J.; Ryder, O. S.; Brady, J. M.; Zhao, D. A Marine Aerosol Reference Tank system as a breaking wave analogue for the production of foam and sea-spray aerosols. Atmos. Meas. Tech. 2013, 6 (4), 1085–1094.

590 591 592

(54)

Holecek, J. C.; Spencer, M. T.; Prather, K. a. Analysis of rainwater samples: Comparison of single particle residues with ambient particle chemistry from the northeast Pacific and Indian oceans. J. Geophys. Res. Atmos. 2007, 112 (22).

593 594 595 596 597

(55)

Martin, A. C.; Cornwell, G. C.; Atwood, S. A.; Moore, K. A.; Rothfuss, N. E.; Taylor, H.; DeMott, P. J.; Kreidenweis, S. M.; Petters, M. D.; Prather, K. A. Transport of pollution to a remote coastal site during gap flow from California’s interior: impacts on aerosol composition, clouds, and radiative balance. Atmos. Chem. Phys. 2017, 17 (2), 1491–1509.

598 599 600

(56) Gard, E.; Mayer, J. E.; Morrical, B. D.; Dienes, T.; Fergenson, D. P.; Prather, K. A. Real-Time Analysis of Individual Atmospheric Aerosol Particles: Design and Performance of a Portable ATOFMS. Anal. Chem. 1997, 69 (20), 4083–4091.

601 602 603

(57)

Noble, C. A.; Prather, K. A. Real-Time Measurement of Correlated Size and Composition Profiles of Individual Atmospheric Aerosol Particles. Environ. Sci. Technol. 1996, 30 (9), 2667–2680.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 30

604 605 606

(58)

607 608 609 610

(59) Cahill, J. F.; Darlington, T. K.; Fitzgerald, C.; Schoepp, N. G.; Beld, J.; Burkart, M. D.; Prather, K. A. Online Analysis of Single Cyanobacteria and Algae Cells under Nitrogen-Limited Conditions Using Aerosol Time-of-Flight Mass Spectrometry. Anal. Chem. 2015, 87 (16), 8039–8046.

611 612 613 614

(60)

Fergenson, D. P.; Pitesky, M. E.; Tobias, H. J.; Steele, P. T.; Czerwieniec, G. a; Russell, S. C.; Lebrilla, C. B.; Horn, J. M.; Coffee, K. R.; Srivastava, A.; et al. Reagentless detection and classification of individual bioaerosol particles in seconds. Anal. Chem. 2004, 76 (2), 373–378.

615 616 617

(61)

Gaston, C. J. Direct Measurements of Marine Aerosols to Examine the Influence of Biological Activity, Anthropogenic Emissions, and Secondary Processing on Particle Chemistry, University of California, San Diego, 2012.

618 619

(62) Russell, S. C. Microorganism characterization by single particle mass spectrometry. Mass Spectrom. Rev. 2009, 28 (2), 376–387.

620 621 622 623

(63)

Gaston, C. J.; Furutani, H.; Guazzotti, S. A.; Coffee, K. R.; Bates, T. S.; Quinn, P. K.; Aluwihare, L. I.; Mitchell, B. G.; Prather, K. A. Unique ocean derived particles serve as a proxy for changes in ocean chemistry. J. Geophys. Res. 2011, 116, D18310.

624 625 626

(64)

Pegus, A.; Kirkwood, D.; Cairns, D. B.; Armes, S. P.; Stace, A. J. Depth profiling of sterically-stabilised polystyrene nanoparticles using laser ablation/ionisation mass spectrometric methods. Phys Chem Chem Phys 2005, 7 (12), 2519–2525.

627 628

(65)

Hardie, L. A.; Eugster, H. P. Evaporation of Seawater: Calculated Mineral Sequences. Science (80-. ). 1980, 208, 498–500.

629 630 631

(66)

Cochran, R. E.; Jayarathne, T.; Stone, E. A.; Grassian, V. H. Selectivity Across the Interface: A Test of Surface Activity in the Composition of Organic-Enriched Aerosols from Bubble Bursting. J. Phys. Chem. Lett. 2016, 7 (9), 1692–1696.

632 633 634 635

(67)

Yoon, Y. J.; Ceburnis, D.; Cavalli, F.; Jourdan, O.; Putaud, J. P.; Facchini, M. C.; Decesari, S.; Fuzzi, S.; Sellegri, K.; Jennings, S. G.; et al. Seasonal characteristics of the physicochemical properties of North Atlantic marine atmospheric aerosols. J. Geophys. Res. 2007, 112 (D4), D04206.

636 637

(68)

Leck, C.; Bigg, E. K. Source and evolution of the marine aerosol—A new perspective. Geophys. Res. Lett. 2005, 32 (19), L19803.

638 639 640 641

(69)

Deng, C.; Brooks, S. D.; Vidaurre, G.; Thornton, D. C. O. Using Raman Microspectroscopy to Determine Chemical Composition and Mixing State of Airborne Marine Aerosols over the Pacific Ocean. Aerosol Sci. Technol. 2014, 48 (2), 193–206.

Sultana, C. M.; Cornwell, G.; Rodriguez, P.; Prather, K. A. FATES: A Flexible Analysis Toolkit for the Exploration of Single Particle Mass Spectrometer Data. Atmos. Meas. Tech. Discuss. 2016, No. October, 1–18.

642

ACS Paragon Plus Environment

24

Page 25 of 30

Environmental Science & Technology

Figure 1

Increasing relative signal from core components

Figure 1. The distribution of the relative sodium ion contribution to the total positive ion intensity for supermicron SSA, generated by frit bubbled seawater analyzed by ATOFMS using five different desorption/ionization laser pulse energies.

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 30

Figure 2

Figure 2. The distribution of total positive ion intensity grouped by the fraction of positive mass spectral sodium intensity for supermicron SSA particles generated by the (a) frit bubbled seawater and (b) MART, and sampled during (c) CIFEX and (d) CalWater-2. Total intensity values were normalized within each dataset.

ACS Paragon Plus Environment

26

Page 27 of 30

Environmental Science & Technology

Figure 3

Figure 3. The sodium ion fraction versus the relative total ion intensity for each positive mass spectrum generated by supermicron SSA from the (a) frit bubbled seawater and (b) MART (1.2 mJ laser pulse energy). Warmer marker shades indicate higher calcium ion fraction in the positive mass spectra.

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 30

Figure 4

Figure 4. (a) The temporal trends for the fraction of ambient particles identified as SSA with sodium fractions of 0.4-1 (Na-Rich) and 0-0.4 (Na-Deficient) shown for two coastal field studies, CalWater2 and CIFEX. (b) Scatter plots of the number fraction of NaDeficient particles against the number fraction of Na-Rich particles and (c) the relative mean signal from nitrate and sulfate markers in the total aerosol population.

ACS Paragon Plus Environment

28

Page 29 of 30

Environmental Science & Technology

Figure 5

Figure 5. A schematic illustrating how two different particle populations could both generate a similar suite of relative mass spectra by laser desorption/ionization, the ionization method utilized by SPMS.

ACS Paragon Plus Environment

29

Environmental Science & Technology

Page 30 of 30

TOC Graphic

ACS Paragon Plus Environment

30